Vous êtes sur la page 1sur 9

J.

of Supercritical Fluids 47 (2009) 373381

Contents lists available at ScienceDirect

The Journal of Supercritical Fluids


journal homepage: www.elsevier.com/locate/supflu

Review

Near critical and supercritical water. Part I. Hydrolytic and hydrothermal


processes
G. Brunner
Institute for Thermal and Separation Processes, Hamburg University of Technology, Eissendorfer Str. 38, D-21073 Hamburg, Germany

a r t i c l e

i n f o

Article history:
Received 21 July 2008
Received in revised form 9 September 2008
Accepted 9 September 2008
Keywords:
Supercritical water
Hydrolysis
Hydrothermal
Biomass
Gasication

a b s t r a c t
The potential of hot and supercritical water in applications to produce useful products, or to process
unwanted compounds into environmentally compatible materials is reviewed. The potential of hot and
supercritical water is high. Water changes its character from a solvent for ionic species at ambient conditions to a solvent for non-ionic species at supercritical conditions. Water at temperatures higher than
ambient boiling temperature can be applied for extraction. At modest temperatures, ionic and polar
species will be extracted. At higher temperatures, in particular approaching the critical temperature,
nonpolar substances are readily dissolved and extracted. Hot pressurized water has a high reactivity. The
reactions are commonly summarized as hydrolysis reactions which are catalyzed by acids, or may arise
from simply hydrothermal transformations. Since CO2 , dissolved in water increases the availability of
protons, the addition of CO2 to liquid water catalyses hydrolysis reactions. Hydrolysis of natural plant
materials provides a route to obtain fuel from non-food plant material. However, difculties associated
with operating conditions have so far limited the large scale implementations.
2008 Elsevier B.V. All rights reserved.

Contents
1.
2.
3.
4.

5.
6.

General introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Properties of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sub-critical water for extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Water for hydrolysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Water in bio-fuel processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.1.
Biomass. Total liquefaction or gasication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.2.
Biomass compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.3.
Starch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.4.
Cellulose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.5.
Sugars, glucose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.6.
Lignin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.7.
Hydrolysis and fermentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
Proteins, amino acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Hydrolysis and hydrothermal reactions in sub- and supercritical water (no oxidative reagent) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions and future development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. General introduction
Interest in the application of hot, pressurized and supercritical
water, started in the late 70s of last century, when it was spurred

Tel.: +49 40428783240.


E-mail address: brunner@tu-harburg.de.
0896-8446/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.supu.2008.09.002

373
374
375
375
375
375
376
376
376
377
377
377
377
378
380
380

by the rst oil crisis, by environmental concerns, and the investigation of supercritical uids. Before that, interest was concentrated
on the properties of high-pressure steam for power plant cycles,
on physicalchemical properties of water, and on hydrothermal
reactions. Renewed interest was concentrated on alternative fuels,
on coal and biomass conversion, and waste disposal. The unique
properties of hot and supercritical water led to the investigation of
specic reactions for production of chemicals, of the formation of

374

G. Brunner / J. of Supercritical Fluids 47 (2009) 373381

particles, and more. The potential of hot and supercritical water


is high and led to extensive research. Operating conditions and
properties of water and species involved cause problems which
are difcult to manage in process equipment which so far led to
only limited practical applications. In the following, after a short
introduction to the properties of water, a review of research efforts
as reected in the publications of the Journal of Supercritical Fluids (JSF), complemented with the personal view of the author is
presented.
2. Properties of water
The specic properties of water for hydrolysis, hydrothermal,
and oxidative processes were pointed out convincingly by M. Modell on several occasions from the 70s of last century on. Water from
ambient to supercritical conditions changes its character from a
solvent for ionic species to a solvent for non-ionic species. Electrochemical properties, e.g. dipole moment decreases from the high
value at ambient conditions, but water in the critical region is
still as polar as acetone. The pH-value decreases by 3 units, providing much more hydronium ions for acid catalyzed reactions.
Just below critical temperature, the ionic product changes tremendously, rendering near critical and supercritical water a much less
polar compound than ambient water. Reactivity of water increases
in the neighborhood of the critical point without as well as with a
catalyst.
It is useful to look at experimental results for the solubility of
SiO2 in water, as determined by mineralogists, who were interested in hydrothermal reactions. This solubility behavior is shown
in Fig. 1. SiO2 is dissolved by water to some extent. This solubility
changes tremendously in the neighborhood of the critical temperature. At relatively low pressures, with increasing temperature, the
solubility drops to practically zero. This effect is one of the major
drawbacks of processing waste, since salts dissolved in the feed
stream will precipitate and eventually block the reactor. But it can
also be seen, that at relatively moderate and high pressures, the
solubility is constant and is even increasing. So far, this aspect has
not been applied to waste treatment.
Among the basic knowledge necessary for a process applying
near critical or supercritical water is the physicalchemical properties of water and its solvent power for organic substances, salts,
and gases. P, v, T and transport properties are that of a supercritical
gas; density is relatively high, up to liquid-like densities, but varies

Fig. 1. Solubility of SiO2 in water [1].

strongly with slight changes in pressure and temperature. Viscosity is of the order of a normal gas and the diffusion coefcient is at
least one order of magnitude higher than that of a liquid. Solubility
of water for gases is high in the critical region. At near critical and
supercritical conditions water and gases like O2 , N2 , NH3 , CO, CO2 ,
are completely miscible. Solvent power of water decreases for inorganic compounds in the critical region. It is drastically reduced in
the region of about 450 C. Organic compounds, on the other hand,
are readily dissolved by water in the near critical and supercritical
region up to total miscibility. But it must be borne in mind that these
statements hold for binary or quasi-binary systems. Phase equilibrium of ternary and multi-component systems of water, organic
compounds, inorganic salts and common gases may deviate from
binary behavior. Properties of water have been investigated intensively for a long time. In our context it may be of interest to look
at phase equilibrium of water and salt components. A small review
has been presented at the 3rd International Symposium on Supercritical Fluids in Strasbourg 1994 [2]. Phase behavior and critical
phenomena for binary mixtures and ternary mixtures of hydrocarbons with water in the critical region have been measured in
particular by Brunner et al. [3], conrming essentially the complete miscibility of water and hydrocarbons in the critical region,
but exhibiting an interesting phase behavior in detail (Fig. 2).
The properties of pure water can be modelled with high accuracy
by thermodynamic models derived from the virial equation of state.
They comprise quite a number of constants and may be impractical for many purposes. Therefore, work has been published on the
modelling of such properties with simpler equations of state such
as the Anderko-Pitzer EoS modied for easier access to parameters, to calculate thermodynamic properties of systems involved in

Fig. 2. (Left): Solubility of a real contamination (extracted from various soil materials) in water. (Right): Course of an extraction of hydrocarbon contaminants from
soil with supercritical water. The full line is obtained by modelling the extraction
[14].

G. Brunner / J. of Supercritical Fluids 47 (2009) 373381

the supercritical water oxidation process [4]. Molecular dynamics


simulations with a simple exible point-charge water model was
applied to model the critical point and the coexistence curve [5].
The local structure around O2 in supercritical water has been investigated using Raman spectroscopy, indicating the absence of large
clustering or depletion of water around O2 under the experimental conditions (380500 C and 5.039.2 MPa). However, depletion
of the local water density around O2 was found in the neighborhood of the critical point [6]. In a similar way, hydrogen bonding
was investigated in methanol/water mixtures [7], indicating that
the methanol molecule, at low methanol contents, is isolated in a
water cage and the structure remains as it is up to 327 C. Electrical
conductance of aqueous solutions of inorganic nitrates was determined with a simple Pt electrode. According to the data, nearly
complete dissociation occurs below 200 C, but appreciable association to monovalent cations occurs at higher temperatures. There
is available a lot of information on water and aqueous systems at
near critical and supercritical temperatures and pressures. It may
prove tedious to look for it, but a good source are the publications
of the IAPWS (International Association for the Properties of Pure
Water and Steam), e.g. [8].
For investigating properties of near critical and supercritical
water, specialised laboratory equipment is necessary. Some of these
installations that have appeared in the JSF include a micro-cell for
potentiometric pH measurements of supercritical aqueous solutions [9], an optical ow cell and apparatus for solubility, salt
deposition and Raman spectroscopic studies in aqueous solutions
near the water critical point [10], a device for direct observation of
channel-tee mixing of high-temperature and high-pressure water
[11], and a reactor for in situ X-ray scattering studies of nanoparticle
formation in supercritical water syntheses [12].

375

[15]. Sub-critical and supercritical water has been also successfully


applied to cleaning of bone materials from organic compounds.
Bones contain lipid and protein compounds, which have to be
removed from the hydroxyapatite, the structural material of bones,
before it can be used for implants. Using hot pressurized water,
lipids and proteins could be extracted. For total extraction of the
proteins, near critical and supercritical water in combination with
a degradation of the proteins proved necessary [16].
4. Water for hydrolysis
4.1. Water in bio-fuel processes
Hot pressurized water has a high reactivity. The reactions are
commonly summarized as hydrolysis reactions. These are reactions in which a compound is split by water according to the formal
reaction:
A B + H OH A H + B OH.

(1)

3. Sub-critical water for extraction

The reaction is catalyzed by acids. Since CO2 , dissolved in water,


increases the availability of protons, the addition of CO2 to liquid water catalyses the hydrolysis reactions. Hydrolysis of natural
plant materials has been intensively investigated, since it provides
a route to obtain ethanol from non-food plant materials such as
straw, wood and bagasse, as summarized by the author in [17,18].
Hydrolysis reactions occurring in hot water are often accompanied
by thermal reactions (pyrolysis). The higher the temperature, the
greater is the extent of thermal reactions. These are hydrothermal
reactions that are discussed below and also in part II of this review.
Reports on investigations of the conversion of bio-polymers seldom distinguish between hydrolytic reactions in hot water and
hydrothermal reactions. Therefore, the classication, into hydrolysis and hydrothermal reactions, as chosen here in this article, is
somewhat arbitrary.

Water at temperatures higher than ambient boiling temperature can be applied for extraction. At lower temperatures, ionic and
polar species will be extracted. At higher temperatures, in particular approaching the critical temperature, nonpolar substances will
be readily dissolved and extracted. Complete dissolution seems not
to be necessary. Water removes nonpolar compounds from substrate by interacting with the substrate and reducing binding forces.
In such a way, hot pressurized, especially near critical water, was
applied to cleaning of soil, removing PAH [13], hydrocarbons and
metals [14].
Cleaning of soil materials contaminated with hydrocarbons was
effectively carried out with sub-critical and supercritical water
in different processing modes. In the semi-continuous xed-bed
extraction and continuous extraction in tubular reactors. Extraction
time needed for cleaning could be reduced drastically, from 6 h for
the semi-continuous extraction to less than 1 min in the continuous
extraction. At a temperature of 380 C and a residence time of 45 s a
98% degree of cleaning could be achieved. But solvent to feed ratio
must be kept high fpr the continuous extraction, >100, corresponding to about 1 wt.% of soil material in the aqueous feed suspension.
Alternatively, a longer tubular reactor (increased residence time)
or a multiple treatment could be applied. Mixed contaminations
of heavy metals and hydrocarbons could be cleaned with water at
temperatures of 250350 C at a pressure of 25 MPa. For destruction, supercritical water oxidation proved to be a good alternative,
as discussed in Part II of the present review appearing also in this
issue of the journal. Polar compounds can be extracted effectively
from plant material with hot pressurized water. One application is
for the extraction of anthraquinones, where anti-oxidative properties of the extract were highest compared to organic extractants

4.1.1. Biomass. Total liquefaction or gasication


Biomass, consisting of plant material and animal products, is
vastly available and either considered as a valuable resource or
as waste. From the part considered as valuable resource, chemical compounds or materials for energy conversion are produced.
The part, considered as waste can be transformed in to disposable
compounds or also in to useful chemical compounds. One example
is the total recycling of organic materials as needed for long-time
secluded missions, e.g. travel to planet Mars. Most of the biological
waste produced can be easily treated by microorganisms, but the
lignocellulosic components cause problems, which can be solved
by hot pressurized water and eventually, for a small residual undissolved part, by application of oxidation in supercritical water, as
investigated in a project sponsored by the European Space Agency
ESA [19]. Another example is the useful transformation of wood
waste. Beside just burning it, it is possible to liquefy wood with hot
pressurized water and produce a combination of fuel gas and fuel
oil [20].
Another route is the gasication of biomass and organic wastes.
Wet biomass and organic wastes can be gasied to produce a
hydrogen rich fuel gas. At 600 C and 250 bar all compounds are
completely gasied by addition of KOH or K2 CO3 , forming a H2
rich gas, containing CO2 as the main carbon compound, with low
concentrations of CO, CH4 and C2 C4 hydrocarbons in the product gas (<1, 3 and <1 vol.%, respectively) [21]. A catalyst seems to
be needed, and catalyst stability and tolerance towards dissolved
inorganics are the main challenges for successful hydrothermal
gasication of wet biomass [22]. The problem is more complex
than can be seen at a rst glance. Beside the necessity to concentrate biomass from a relatively large area and transport it to

376

G. Brunner / J. of Supercritical Fluids 47 (2009) 373381

uation to food. Cellulose and hemi-cellulose, being available from


agricultural sources without competition to food, can be transformed to mono-sugars, but so far scientic and technological
development has not led to a commercial process, which has been
installed. Nevertheless, this will be the future for bio-ethanol production.

4.1.3. Starch
Hydrolysis of starch to mono-sugars is an old industrial process. The hydrolysis reaction was catalyzed with mineral acids, and
is now replaced by a bio-transformation. Microorganisms for this
reaction are selective but slow. Hydrolysis using CO2 , dissolved in
hot pressurized water, without mineral acids was investigated to
evaluate the feasibility of such a process [24]. From corn starch at
concentrations from 0.2 up to 10 wt.% in the aqueous feed slurry and
pressures of 60240 bar, temperatures of 170380 C in a tubular
reactor with residence times of about 180 s, high yields of glucose
can be obtained with dissolved and dissociated carbon dioxide as
a catalyst. There are only few by-products, and their concentration
can be kept low by short residence time. Fig. 3 shows the results
of the hydrolysis of starch at 230 C and 240 bar, for an example.
Glucose yield is increased from 5% to 60% by adding carbon dioxide.

Fig. 3. Product yields at 230 C and 240 bar (left); inuence of carbon dioxide on
glucose yield (right). CO2 concentrations are in percentage of saturation.

a processing facility, the technological problems are multiple due


to the non-controllable composition of the biomass feed, in particular with respect to inorganic contaminations. More promising
is a route, where the biomass is liqueed to an oil-product, the
oil is then transported to a central processing facility, where an
effective gasication to synthesis gas is carried out, from which,
with the well known FischerTropsch-process, fuel hydrocarbons
can be produced [23].
4.1.2. Biomass compounds
Plant biomass consists of a number of compounds which can be
processed by pressurized hot water, like sugars, starch, cellulose,
hemi-cellulose, and lignin. Starch, cellulose, and hemi-cellulose are
sugar-polymers and can be transformed to sugar-monomers. From
starch, mono-sugars are produced straightforward by microorganisms or by hydrolysis and can be further transformed to ethanol,
which is of some concern nowadays because of the competitive sit-

4.1.4. Cellulose
Cellulose, the major component of plant biomass reacts in
hot compressed water to oligomer sugars, monomer sugars,
and various degradation products as pyruvaldehyde and hydroxymethylfurfural (HMF), as e.g. reported by Sasaki et al. [25]. The
decomposition of cellulose can lead to a reaction mixture of oil,
gases, residue and an aqueous phase, as reported by Minova et al.
[26]. But the batch reactor used and the residence times in the order
of 30 min to several hours are no basis for a process. According to
all reports, hydrolysis can play an important role in forming glucose and oligomers, which can decompose quickly to non-glucose
aqueous products, oil, char and gases. An alkali catalyst, inhibits the
formation of char from oil and stabilizes the oil, while a nickel catalyst catalyzes the steam reforming reaction of aqueous products
as intermediates and the reaction to methane. The formation of
methane is the preferred reaction pathway at lower temperatures,
while the production of hydrogen proceeds in the supercritical
region. A substantial conversion to gaseous species is, however, usually accomplished at much higher temperatures (400500 C) and
residence times in the order of hours. This observation was supported by our own measurements at temperatures of 350375 C
and residence times from 20 to 40 s, which yielded a contribution
of gas species to the carbon balance of only about 1 wt.%. A marked
gasication of cellulose can therefore be excluded at the conditions
of hot pressurized water hydrolysis [19].

Fig. 4. Degree of liquefaction f versus residence time  at P = 250 bar.

G. Brunner / J. of Supercritical Fluids 47 (2009) 373381

The overall degree of conversion (liquefaction) of microcrystalline cellulose was determined in a plug-ow type reactor.
The operating pressure had a negligible inuence on the degree
of liquefaction, but temperature markedly affects the rate of reaction as illustrated in Fig. 4. The degree of liquefaction f is the ratio
of dissolved carbon in the efuent to the total carbon of the feed
suspension.
ln(1 f ) = k 

(2)

where k is the reaction rate constant of the cellulose hydrolysis.


The rate of liquefaction increases with temperature, leading to a
complete conversion to soluble products in less than 30 s at 310 C.
A further increase in temperature results in an even more rapid
degradation, with a complete conversion within seconds.
The addition of carbon dioxide and the resulting decrease in
pH leads to an increased rate of liquefaction at lower temperature (250 C) compared to pure water. Glucose yield during the
course of the reaction is shown in Fig. 5. Rate constants for formation of glucose are higher for starch than for cellulose. Glucose
is not stable at reaction conditions and is degraded by consecutive
reactions to different products, including pyruvaldehyde, levoglucosan, and hydroxymethylfurfural. Reaction kinetics are discussed
in a separate paper [4].
The optimum glucose yields are obtained with residence times
in the range of seconds to minutes but remain relatively low (e.g.
compared to the values for hydrolysis of starch). The maximum
yield of glucose increases with temperature, at shorter residence
times. Around the critical point, the hydrolysis rate jumps to more
than an order of magnitude higher level and becomes faster than
the glucose or oligomer decomposition rate [25]. The reported glucose yield reached a maximum of 50% at 400 C and 25 MPa, at a
residence time of 0.00025 s.
4.1.5. Sugars, glucose
Reaction of glucose and fructose in water at supercritical conditions resulted in a number of compounds. Dehydration of d-glucose
in high-temperature water at pressures up to 80 MPa yielded
5-hydroxymethylfurfural (5-HMF), 1,2,4-benzenetriol (BTO) and
furfural. Dehydration reaction to 5-HMF and hydrolysis of 5HMF were both enhanced by the increase in water density at
400 C [27]. An analogue study of reactions of d-fructose in
water yielded at high-temperature (400 C) and moderate pressure
(40 MPa) conditions products as glyceraldehyde, dihydroxyacetone
and pyruvaldehyde. High-temperature, high-pressure conditions

Fig. 5. Glucose yield from sub-critical water hydrolysis of cellulose at different residence times and temperatures; and the effect of dissolved CO2 .

377

enhanced retro-aldol reactions and water related reactions such as


hydrolysis and dehydration. Comparison between d-fructose and
d-glucose experiments showed that higher yields for 5-HMF were
obtained starting from d-fructose than d-glucose. Higher yields
of furfural were obtained starting from d-glucose than from dfructose [28].
4.1.6. Lignin
Lignin is also a major component of plant biomass. To make
available the elemental chemical compounds of lignin for further
application is still a task for research. During studies on the liquefaction of biomass, it was found that the biomass could be liqueed
by hydrolysis up to 7080% [19]. The efuents were subsequently
treated by biological degradation. Overall efciency of COD removal
increased to 9095%. No toxic effects on the microorganisms were
observed due to the prior hydrolytic treatment. The remaining
compounds were attributed to lignin-derivates. They could not be
reacted with water. Oxidation in near critical water by hydrogen
peroxide converted all solid material, mostly to gaseous products.
Only about 10% of the initial carbon load remained in the aqueous
phase, with the main product being acetic acid [19].
4.1.7. Hydrolysis and fermentation
To enhance yield in mono-sugars for the treatment of biomass,
the combination of hot pressurized water hydrolysis and fermentation was considered. As an example, rice bran, milled and
defatted, consisting of starch, 27% cellulose, 37% hemi-cellulose,
5% lignin was selected [17]. After hydrolysis, the reaction products,
mainly oligomer sugars, are easily transformed to mono-sugars by
enzymes. Hot water hydrolysis at 200 C resulted in 5% glucose
and 35% xylose. Hot water hydrolysis with subsequent enzymatic
treatment resulted in 70% glucose and 70% xylose, relative to
the maximum obtainable content in the feed. Results leading in
a comparable direction, were obtained on the hydrolysis of carboxymethylcellulose. Cellulase, immobilized on silica gel was used
as biocatalyst for hydrolysis in a high-pressure system of CO2 /H2 O.
The residual activity of the immobilized cellulase at 110 C was still
high and it could be reused for more than 20 times without any
signicant loss of activity [29].
4.2. Proteins, amino acids
Proteins are the other type of important bio-polymers. The reaction of hot pressurized water can be of interest for producing
oligomers and amino acids as the building blocks of the proteins. In
the hydrolysis reaction for proteins, rst a proton is attached to the
nitrogen atom of the peptide bonding. This leads to a splitting of the
bonding, forming a carbo-cation and an amino group. In the next
step, a hydroxide ion, from a dissociated water molecule, attaches
to the carbon-cation, forming a carboxy group.

Hydrolysis of a model protein (BSA, bovine serum albumin) and


sklero-proteins like feathers and hair, carried out in a continuous
plug-ow reactor, resulted in a total liquefaction of the proteins
and in the formation of amino acids [30]. Production of amino
acids depends mainly on reaction temperature, with an optimum
at 310 C. Pressure in the range of 1527 MPa had no signicant
effect on the reaction. At 250 C the amino acid yield increases

378

G. Brunner / J. of Supercritical Fluids 47 (2009) 373381

up to a residence time of 300 s and then decreases due to decomposition reactions. Considerable quantities of glycine and alanine
were produced from decomposition of complex amino acids. Other
amino acids were only found in traces. Addition of carbon dioxide resulted in higher yields due to acid hydrolysis of the peptide
bonds. At 250 C and 25 MPa an amino acid yield of 150.3 mg/(g
BSA) was obtained by sub-critical water being saturated with CO2
to approximately 90%. The experiments with duck feathers (without the addition of CO2 ) led to amino acid yields (122.0 mg/(g
DFK) at 900 s) higher than for BSA without addition of CO2 due
to the shorter chain length of the sklero-keratin molecule. Thus,
sub-critical water hydrolysis can be an efcient process for recovering amino acids from organic protein-rich waste-materials, such
as hairs and feathers [31].
Hydrolysis kinetics of starch, cellulose (polysaccharides), and
proteins (polypeptides) could be modelled by a single consecutive reaction following rst order kinetics. Rate constants of the
hydrolytic conversion to the resulting monomers (glucose and
amino acids), strongly depend on the type of bond. Peptide bonds in
proteins are much more stable than the -1,4- and -1,6-glycosidic
linkages in cellulose and starch, respectively. The stability of the
resulting monomers and their conversion to further degradation
products were determined. The addition of carbon dioxide to water
up to about 250 C resulted in a signicant increase of reaction rates
[18].
5. Hydrolysis and hydrothermal reactions in sub- and
supercritical water (no oxidative reagent)
Quite a number of investigations have concentrated on
hydrothermal and hydrolysis reactions. Hydrolytic and hydrothermal reaction are not and cannot always be separated. What is a
hydrothermal reaction? The term hydrothermal was used in the
context of synthesis of minerals in natural deposits by crystallization from hot pressurized aqueous solutions.
Hydrolysis in supercritical water is determined by the extreme
pressure dependence of the solvent propertiesdensity, dielectric
constant, and solubility parameter. Klein et al. [32] investigated the
reaction mechanism in that region. Molecules containing a saturated carbon attached to a heteroatom-containing leaving group
undergo parallel pyrolysis and hydrolysis reactions in supercritical
water. Selectivity and rate constant to hydrolysis increased with
water density and with the addition of salts. At low salt concentration, the addition of salts to the reaction mixture increased the
hydrolysis rate, while having no effect on a background pyrolysis
reaction rate [33]. This is consistent with a polar hydrolysis reaction mechanism wherein the rate constant would be increased with
increases in the solvent polarity. At higher salt concentrations, the
hydrolysis rate reached a maximum then decreased, approaching
the rate observed in the absence of salts, which corresponds to
solution phase behavior. Behavior of various compounds at hydrolysis conditions and at hydrothermal conditions have been reported.
Those published in the Journal of Supercritical Fluids are shortly
reviewed below.
Behavior of acids and bases in the neighborhood of the critical
point is of major interest. Johnston and Chlistunoff [34] investigated
the neutralization of acids and bases in sub-critical and supercritical water using KOHacetic acid or NH3 acetic acid systems as
examples. From 25 C to the critical temperature of water, the dissociation constant for HCl decreases by 13 orders of magnitude,
and thus, the basicity of Cl becomes signicant. Consequently, the
addition of NaCl to HCl raises the pH.
The group of Antal [35] investigated mechanism and kinetics of the acid-catalyzed dehydration of ethanol in supercritical
water. In the presence of a low concentration (<0.01 mol dm3 ) of

sulfuric acid, ethanol undergoes rapid and selective dehydration


to ethene in supercritical water. The kinetics of this reaction are
consistent with an acid-catalyzed E2 mechanism. In the group of
Tester [36] oxidation and hydrolysis of ethanol were investigated in
supercritical water in a plug-ow reactor system. In the hydrolysis
experiments, ethanol did not react to a signicant degree relative
to the conversions observed in oxidation experiments. In the oxidation experiments, oxygen concentration was set for complete
mineralization. The major products of the reaction were acetaldehyde and formaldehyde in the liquid phase and carbon monoxide
and carbon dioxide in the gas phase.
The dehydration of several biomass-derived polyols in
sub- and supercritical water resulted in main products of
1,4-anhydroerythritol and propionaldehyde [37]. For catalytic performance zinc, nickel, copper, magnesium and sodium sulfate as
well as sulfuric acid were tested, with a positive catalytic effect for
bivalent transition metal salts.
Decomposition of glycerol in near critical and supercritical
water was investigated in a plug-ow reactor at 349475 C,
2545 MPa, and reaction times from 32 to l65 s at different initial
concentrations [38]. Conversion between 0.4 and 31% was observed.
The main products of the glycerol degradation are methanol,
acetaldehyde, propionaldehyde, acrolein, allyl alcohol, ethanol,
formaldehyde, carbon monoxide, carbon dioxide, and hydrogen.
The non-Arrhenius behavior of the overall degradation, the pressure dependence of the reaction rate, and product distribution
indicates to two competing reaction pathways, one of ionic reactions, preferred at higher pressures and lower temperatures, the
other a free radical degradation, which dominates at lower pressures and higher temperatures.
For ester hydrolysis, ethyl acetate was studied as a model compound in the pressure and temperature range of 2330 MPa and
250400 C. Based on this results and on the hydrolysis of other
esters, the reaction mechanisms in sub- and supercritical water
are discussed [39]. The ester hydrolysis proceeds selectively to the
expected acid and alcohol without involving acids and bases as
catalysts. In the sub-critical region an Aac 2-mechanism is found.
Above the critical temperature of water, the reaction mechanism is
the direct nucleophilic attack of water. The two concurrent mechanisms lead do the same products.
Hydrolysis of triacylglycerides was studied for canola oil [40]
for formation of free fatty acids (FFA) by oil hydrolysis in combination with supercritical CO2 and N2 . Reactions with CO2 were
conducted at 250 C, 1030 MPa, and using 1:3, 1:17 and 1:70 canola
oil to water initial molar ratio (o/w) in a batch reactor system. The
maximum rate of FFA production was not affected by pressure or
added supercritical medium, but was delayed at 30 MPa. Hydrolysis
increased signicantly as the amount of water was increased from
1:3 to 1:17 and 1:70 o/w.
Reactions of diphenylether in supercritical water were reported
by Penninger et al. [41]. At low water density, from 0 to approximately 0.3 g/cm3 , the conversion of diphenylether decreases.
Products are typical for radical-type polycondensation reactions.
At water densities greater than 0.4 g/cm3 these products vanish,
the conversion of diphenylether increases again and forms phenol
as sole reaction product. This indicates ionic hydrolysis as the governing chemistry. Dilute solutions of NaCl in supercritical water,
as also reported by Penninger et al. [42], in concentrations from 0
to 3.1 wt.% have unexpected inuence on the rate of hydrolysis. At
430 C and a density of 0.46 g/cm3 , the rate decreased with addition
of small amounts of salt, down from the value at zero salt. The rate
increased with addition of more salt (3.1 wt.%) to almost twice that
of the zero salt rate. The decrease of the rate is attributed to the
formation of the ion pairs H+ Cl ; the excess of Cl ions that prevail
from ionic (partial) dissociation of NaCl and capture the protons

G. Brunner / J. of Supercritical Fluids 47 (2009) 373381

generated by self-dissociation of water. At higher salt content the


mechanism is rationalized with Lewis acid/base behavior of Na+
and Cl ions in solution. The rate of hydrolysis is proportional to
the square root of the salt concentration in the supercritical water
reaction medium.
For conversion of carbohydrates it was found by Bicker et al.
[43] that small amounts of a salt can increase the yield of lactic acid drastically. The group of Vogel [44] also investigated the
dehydration of d-fructose to hydroxymethylfurfural. In supercritical water only unsatisfying yields of HMF are achieved by the
dehydration of d-fructose. But in the reaction in sub- and supercritical methanol the resulting 5-methoxymethylfurfural was obtained
with 79% selectivity at 99% conversion and with sub-critical acetic
acid 5-acetoxymethylfurfural (MMF) was obtained with 38% selectivity at 98% conversion, similar to previously obtained results for
d-fructose to HMF in sub- and supercritical acetone/water mixtures. The authors propose a continuous production process for
HMF and MMF.
Hydrolysis of nitriles was investigated by the group of Klein
[45,46]. The reactions of butyronitrile were investigated in hightemperature water at 330 C and various pressures ranging from
128 to 2600 bar [45]. Residence times ranged from 5 to 180 min. The
product spectrum included butanamide, butyric acid and ammonia; gas formation was negligible. A four-step autocatalytic rate
model incorporating the product acid as the catalytic species tted the data well. The kinetics of benzonitrile hydrolysis were used
to investigate reactor wall effects [46]. Repeated reactor use had
no effect on either the rate constant or product selectivity as well
as the use of stainless steel and titanium reactors. Wall inuence
was within the experimental uncertainty for the homogeneous rate
constant (6.1%). From that is concluded that the laboratory-deduced
rate constants can be applied to kinetics at commercial conditions.
Product distribution and reaction pathways for methylene chloride hydrolysis and oxidation was investigated by Testers group
[47]. Reactions were carried out with dilute feeds at 24.6 MPa,
25600 C at residence times of 7 to 23 s in a tubular reactor system. The products detected were formaldehyde, hydrochloric acid,
carbon monoxide, hydrogen, methanol, and carbon dioxide, with
trace amounts of other compounds. The main route for CH2 Cl2
breakdown was via sub-critical hydrolysis to formaldehyde and
HCl, followed by decomposition of formaldehyde to CO and H2 ,
and subsequent CO conversion to CO2 and H2 . At 600 C and 6 s
residence time under oxidation conditions, CO2 and HCl were the
only products at complete destruction of all compounds (>99.99%
of total carbon). An analysis of hydrolysis kinetics, phase equilibria,
and modelling has been reported also [48].
Properties and synthesis reactions in hot compressed water
as reaction medium and reactant have been reviewed by Kruse
and Dinjus [4051]. Macroscopic and microscopic properties of
hot pressurized water are described, and a summary of published
synthesis reactions is presented. For the thermal degradation of
tert-butylbenzene in hot pressurized water, a high-pressure reaction model was developed. The model consists of 171 elementary
reactions. Simulation and experimental results correlated reasonably. The reaction in hot pressurized water is strongly slowed down
by a factor of 1000 in comparison to the reaction at low pressure
in an inert environment. It is assumed that a cage effect of water
molecules reduces the reactivity of these species. Detailed reaction
mechanisms are presented. For application, methods for determining kinetic parameters for simplied hydrothermal oxidation
reactions are of use [52]. Three different methods, namely pseudorst-order kinetics, multiple linear regression and RungeKutta
algorithm, were used to determine the kinetic parameters, with
similar results. The RungeKutta algorithm was found to be more
convenient for the kinetic parameter determination.

379

Formaldehyde (HCHO) reaction in supercritical water was


studied with batch experiments to determine the water density
dependence [53]. At higher water densities, CH3 OH yields were
high, at low water densities, CO yields increased.
Reaction mechanisms for the decomposition of phenanthrene
and naphthalene under hydrothermal conditions have been
investigated by Onwudili and Williams [54]. Polycyclic aromatic
compounds have been oxidized in a hydrothermal oxidation batch
reactor. At lower temperatures the polycyclic aromatic compounds
were thermally cracked but as the temperature increased, hydroxylation of the aromatic moiety occurred leading to a series of
activities including ring-opening and rearrangement reactions. It
was found that up to 99 wt.% destruction of the polycyclic aromatic
compounds occurred at supercritical conditions.
Hydrothermal synthesis of low-molecular uorescent organic
molecules derived from an amino acid, glycine, using a ow reactor with rapid expansion cooling is reported by Futamura et al. [55].
The product mixture solution had emission from blue to ultraviolet.
HPLC fractions of the products also had uorescence and included
CSH6 N2 O and its methyl derivative. The uorescent compounds
suggest carboncarbon bond formation as well as dehydration condensation.
The decomposition in sub- and supercritical water of the amino
acids alanine and glycine was studied [56]. The main reactions
were amino acid decarboxylation and amino acid deamination,
which could be modelled by simple global rate laws.
The reactivity of methylamine in supercritical water was studied in a batch reactor at temperatures between 386 and 500 C
[57]. The major products measured are ammonia and methanol.
For water densities less than 0.28 g/cm3 and pressures less than
to 25 MPa, the effect of water on the reaction rate appears to be
negligible, and there is little evidence of hydrolysis. Under these
conditions, the reaction seems to be governed by pyrolysis. At
higher water densities, hydrolysis becomes more important and
the methanol yield increases with water density.
The effect of water density on the rate constant for the decomposition of aliphatic nitrocompounds, was investigated by Anikeev
et al. [58]. The density dependence of the rate constant of decomposition for each nitromethane, nitroethane, and 1-nitropropane,
correlates with the density dependence of the H3 0+ concentration
in dissociated supercritical water.
Thiodiglycol [(HOC2 H4 )2 S] hydrolysis and oxidation in sub-,
and supercritical water was studied by Testers group [59]. Under
supercritical conditions (T = 400525 C and P = 25 MPa), thiodiglycol degradation occurred rapidly with and without oxidant. At
sub-critical temperatures from 100 to 360 C, hydrolysis reaction
products identied included carbon monoxide, carbon dioxide,
hydrogen, methane, ethylene, acetaldehyde, acetic acid, formic
acid, thioxane, sulfuric acid, hydrogen sulde, and elemental sulfur.
Hydrothermal stability of six aromatic carboxylic acids has
been investigated by Savage [60]. Benzoic acid was the most
stable, showing negligible degradation after 1 h of hydrothermal
treatment at 350 C. Terephthalic acid, 2,6-naphthalene dicarboxylic acid and isophthalic acid were stable after 1 h at 300 C,
but they decarboxylated to form monoacids in 1015% yields at
350 C. Trimellitic anhydride decomposed completely after 30 min
at 350 C, but showed no appreciable decomposition after 30 min
at 250 C. Terephthalic acid and isophthalic acid were the main
degradation products, but o-phthalic acid was also formed in small
amounts at 350 C. The o-phthalic acid conversion to benzoic acid
was 73% after 60 min at 300 C, but the diacid remained stable at
250 C for 1 h.
Cyanamide, dicyandiamide and related cyclic azines were
reacted in water at 100300 C in a sealed 316 SS tube (27.5 MPa)
for the purpose of characterizing the hydrothermolysis chemistry

380

G. Brunner / J. of Supercritical Fluids 47 (2009) 373381

of cyanamide [61]. The conversion of cyanamide to dicyandiamide dominates at 100175 C. At 175250 C, for reaction times
shorter than 15 min, the major pathway is hydrolysis of the
cyanamidedicyandiamide mixture to CO2 and NH3 . Above about
225 C, hydrolysis of these cyclic azines to aqueous NH3 and CO2
occurs. At 300 C the conversion of all compounds to CO2 and NH3
is complete in 10 min.
Of major interest for hydrolysis and hydrothermal reactions are
polymers in particular for the treatment of waste. Waste polyvinyl
chloride plastics were treated at hydrothermal conditions [62].
Chlorine in PVC dissolved in water as hydrochloric acid. No harmful chlorinated organic compounds were observed in the liquid and
gas fractions after treatment at 300 C. Between 250 and 350 C, this
technique produced polyene as a residual solid, and low-molecular
weight aromatic and aliphatic compounds in the liquid and gas
fractions. Further decomposition at over 350 C in supercritical
water produced acetone, phenol, benzene, benzene derivatives, and
aliphatic alkane and alkene in the liquid and gas fractions. The
combustion enthalpy of the residual solid was 9270 kcal/kg.
Polyethylene terephthalate was reacted in water at pressures
up to 173 MPa and temperatures up to 490 C [63]. Over the
range of PET concentrations studied (1259 wt.%), most systems
became homogeneous. For the case where samples were heated
slowly, the homogenization temperature was around the PET melting point (241 C). For PET samples that were rapidly heated and
underwent a solidliquid transition upon heating, the homogeneous temperature was between 297 and 318 C. For samples that
were rapidly heated and underwent crystallization during heating, the homogenization temperature was between 360 and 396 C.
The homogenization temperature for terephthalic acid + water was
found to be around 356 C.
In a similar way, polyethylene phase behavior and reaction in
supercritical water at pressures up to 2.6 GPa and temperatures up
to 670 C have been investigated [64]. When PE + water (1230%
PE) mixtures were rapidly heated at initial pressures ranging from
110 to 690 MPa, PE rst melted and formed a liquid spherule PE
phase. The results of this study show that PE and water remain
as a heterogeneous system over the polymer (1230% PE) compositions studied during heating and reaction in supercritical water.
Only after PE decomposes to lower molecular weight hydrocarbons,
above about 565 C, homogeneous reaction conditions result.
At more accessible conditions, an investigation on pyrolysis
of polyethylene and n-hexadecane [65] was carried out in a
batch reactor at temperatures from 400 to 450 C, a reaction
time of 30 min, and water density between 0 and 0.42 g/cm3 .
In supercritical water, higher yields of shorter chain hydrocarbons, higher l-alkene/n-alkane ratio, and higher conversion were
obtained than from pyrolysis in Ar. Pyrolysis rate of nC16 in supercritical water was almost the same as that in 0.1 MPa argon (Ar)
atmosphere.

6. Conclusions and future development


Properties of sub- and supercritical water are well known, also
in combination with gases, salts, and many organic compounds.
The knowledge of these properties is a prerequisite for carrying out
processes under these conditions. The special properties of water
can be used for extraction of polar compounds, for the hydrolytic
transformation of monomers and polymers. Catalysts may be useful
in special cases. In particular, CO2 , dissolved in liquid hot pressurized water may be of advantage for producing higher yields
of product components at milder conditions, avoiding formation
of by-products. Hydrolysis and hydrothermal reactions of organic
molecules lead to a bunch of compounds, and it proves difcult to

obtain special compounds at high concentrations. But, some investigations have shown, that with optimized conditions, it may be
possible. A combination of hydrolytic and bio-catalytic processes
seems to be very promising for processing biomass. For the future,
we will see the application of hot and pressurized water to biomass
conversion, the application to waste treatment, and in some favorable cases to production of single product components.
References
[1] G. Brunner. Gas extraction. An Introduction to Fundamentals of Supercritical
Fluids and the Application to Separation Processes. Steinkopff, Springer, Darmstadt, New York, 1994. ISBN 3-7985-0944-1; ISBN 0-387-91477-3, p. 62 (Data
from: G.C. Kennedy, A portion of the system silicawater. Econ. Geol. 45 (1950)
629653).
[2] A. Firus, G. Brunner, Decontamination of soil and oxidation of the extracts by/in
supercritical water: the underlying phase equilibria of binary and multicomponent systems. Proceedings of Third International Symposium on Supercritical
Fluids, M. Perrut, G. Brunner (eds.), Strasbourg 1994, ISBN 2-905267-23-8, vol.
3, pp. 195200.
[3] E. Brunner, M.C. Thies, G.M. Schneider, Fluid mixtures at high pressures: phase
behavior and critical phenomena for binary mixtures of water with aromatic
hydrocarbons, J. Supercrit. Fluids 39 (2006) 160173.
[4] M.D. Bermejo, A. Martin, M.J. Cocero, Application of the Anderko-Pitzer EoS
to the calculation of thermodynamical properties of systems involved in the
supercritical water oxidation process, J. Supercrit. Fluids 42 (2007) 2735.
[5] T.I. Mizan, Ph.E. Savage, R.M. Ziff, Critical point and coexistence curve for a exible, simple point-charge water model, J. Supercrit. Fluids 10 (1997) 119125.
[6] K. Sugimoto, H. Fujiwaral, S. Koda, Raman spectroscopic study on the local structure around O2 in supercritical water, J. Supercrit. Fluids 32 (2004) 290302.
[7] T. Ebukuro, A. Takami, Y. Oshima, S. Koda, Raman spectroscopic studies on
hydrogen bonding in methanol and methanol/water mixtures under high temperature and pressure, J. Supercrit. Fluids 15 (1999) 7378.
[8] Physical chemistry of aqueous systems: meeting the needs of industry, in: H.J.
White Jr., J.V. Sengers, D.B. Neumann, J.C. Bellows (Eds.), Proceedings of International Conference on the Properties of Steam (12th, 1994, Orlando, FL), Begell
House, New York, 1995.
[9] K. Sue, F. Ouchi, K. Arai, Microcell for potentiometric pH measurements of
supercritical aqueous solutions, J. Supercrit. Fluids 39 (2006) 271276.
[10] W.S. Hurst, M.S. Hodes, W.J. Bowers Jr., V.E. Bean, J.E. Maslar, P. Grifth, K.A.
Smith, Optical ow cell and apparatus for solubility, salt deposition and Raman
spectroscopic studies in aqueous solutions near the water critical point, J.
Supercrit. Fluids 22 (2002) 157166.
[11] T. Aizawa, Y. Masuda, K. Minami, M. Kanakubo, H. Nanjo, R.L. Smith, Direct
observation of channel-tee mixing of high-temperature and high-pressure
water, J. Supercrit. Fluids 43 (2007) 222227.
[12] M. Bremholm, H. Jensen, S. Brummerstedt Iversen, B. Brummerstedt Iversen,
Reactor design for in situ X-ray scattering studies of nanoparticle formation in
super-critical water syntheses, J. Supercrit. Fluids 44 (2008) 385390.
[13] J. Kronholm, J. Kalpala, K. Hartonen, M.-L. Riekkola, Pressurized hot water
extraction coupled with supercritical water oxidation in remediation of sand
and soil containing PAHs, J. Supercrit. Fluids 23 (2002) 123134.
[14] G. Brunner, B. Misch, A. Firus, K. Nowak, Cleaning of soil with supercritical water
and supercritical carbon dioxide, in: R. Stegmann, G. Brunner, W. Calmano, G.
Matz (Eds.), Treatment of SoilFundamentals, Analysis, Applications, Springer,
Berlin, 2001, pp. 491517.
[15] B. Pongnaravane, M. Goto, M. Sasaki, T. Anekpankul, P. Pavasant, A. Shotipruk,
Extraction of anthraquinones from roots of Morinda citrifolia by pressurized hot
water: antioxidant activity of extracts, J. Supercrit. Fluids 37 (2006) 390396.
[16] D. Doncheva, G. Brunner, Cleaning of animal-derived bone material for
implantation by combined extraction/reaction process of organic matrix with
subcritical water and characterisation of hydrolysates, in: Proceedings of Fifth
International Symposium on High Pressure Process Technology and Chemical
Engineering, Segovia, Spain, June 2427, 2007.
[17] G. Brunner, From plant materials to ethanol by means of supercritical uid
technology, in: Proceedings of Fifth International Symposium on High Pressure Process Technology and Chemical Engineering, Segovia, Spain, June 2427,
2007.
[18] T. Rogalinski, K. Liu, T. Albrecht, G. Brunner, Hydrolysis kinetics of biopolymers
in subcritical water, in: Proceedings of Fifth International Symposium on High
Pressure Process Technology and Chemical Engineering, Segovia, Spain, June
2427, 2007.
[19] G. Lissens, W. Verstraete, T. Albrecht, G. Brunner, C. Creuly, J. Seon, G. Dussap, Ch.
Lasseur, Advanced anaerobic bioconversion of lignocellulosic waste for bioregenerative life support following thermal water treatment and biodegradation
by Fibrobacter succinogenes, Biodegradation 15 (2004) 173183.
[20] Th. Willner, G. Brunner, Umwandlung von Holz unter dem Einu von Wasserstoff und Wasser unter hheren Drcken, Chemie-Ing. -Technol. 66 (1994)
7274.
[21] H. Schmieder, J. Abeln, N. Boukis, E. Dinjus, A. Kruse, M. Kluth, G. Petrich, E.
Sadri, M. Schacht, Hydrothermal gasication of biomass and organic wastes, J.
Supercrit. Fluids 17 (2000) 145153.

G. Brunner / J. of Supercritical Fluids 47 (2009) 373381


[22] M.H. Waldner, F. Krumeich, F. Vogel, Synthetic natural gas by hydrothermal
gasication of biomass. Selection procedure towards a stable catalyst and its
sodium sulfate tolerance, J. Supercrit. Fluids 43 (2007) 91105.
[23] L. Leible, S. Klber, G. Kappler, S. Lange, E. Nieke, P. Proplesch, D. Wintzer, B.
Frni, Kraftstoff, Strom und Wrme aus Stroh und Waldrestholz Eine systemanalytische Untersuchung, Forschungszentrum Karlsruhe, Karlsruhe, 2007
(Wissenschaftliche Berichte, FZKA 7170).
[24] K. Liu, G. Brunner, Hydrolysis of some biopolymers with water and carbon
dioxide at high pressures and temperatures, in: Proceedings of International
Meeting of the GVC-Fachausschu Hochdruckverfahrenstechnik, Karlsruhe,
Germany, March 35, 1999, pp. 123126 (Also in: Wissenschaftliche Berichte
FZKA 6271).
[25] M. Sasaki, B. Kabyemela, R. Malaluan, S. Hirose, N. Takeda, T. Adschiri, K. Arai,
Cellulose hydrolysis in subcritical and supercritical water, J. Supercrit. Fluids 13
(1998) 261268.
[26] T. Minowa, F. Zhen, T. Ogi, Cellulose decomposition in hot-compressed water
with alkali or nickel catalyst, J. Supercrit. Fluids 13 (1998) 253259.
[27] T.M. Aida, Y. Sato, M. Watanabe, K. Tajima, T. Nonaka, H. Hattori, K. Arai, Dehydration of -glucose in high temperature water at pressures up to 80 MPa, J.
Supercrit. Fluids 40 (2007) 381388.
[28] T.M. Aida, K. Tajima, M. Watanabe, Y. Saito, K. Kuroda, T. Nonaka, H. Hattori, R.L.
Smith Jr., K. Arai, Reactions of d-fructose in water at temperatures up to 400 C
and pressures up to 100 MPa, J. Supercrit. Fluids 42 (2007) 110119.
[29] M. Paljevac, M. Primozic, M. Habulin, Z. Noyak, Z. Knez, Hydrolysis of carboxymethyl cellulose catalyzed by cellulose immobilized on silica gels at low
and high pressures, J. Supercrit. Fluids 43 (2007) 7480.
[30] T. Rogalinski, S. Herrmann, G. Brunner, Production of amino acids from bovine
serum albumin by continuous sub-critical water hydrolysis, J. Supercrit. Fluids
36 (2005) 4958.
[31] T. Rogalinski, G. Brunner, Production of amino acids from bovine serum albumin and duck feather keratin by continuous sub-critical water hydrolysis, in:
Proceedings of Seventh World Congress of Chemical Engineering, Glasgow,
Scotland, July, 2005.
[32] M.T. Klein, L.A. Torry, B.C. Wu, S.O. Townsend, S.C. Paspek, Hydrolysis in supercritical water: Solvent effects as a probe of the reaction mechanism, J. Supercrit.
Fluids 3 (1990) 222227.
[33] L.A. Torry, R. Kaminsky, M.T. Klein, M.R. Klotz, The effect of salts on hydrolysis
in supercritical and near-critical water: reactivity and availability, J. Supercrit.
Fluids 5 (1992) 163168.
[34] K.P. Johnston, J.B. Chlistunoff, Neutralization of acids and bases in subcritical and supercritical water: acetic acid and HCl, J. Supercrit. Fluids 12 (1998)
155164.
[35] X. Xu, C. De Almeida, M.J. Antal Jr., Mechanism and kinetics of the acid-catalyzed
dehydration of ethanol in supercritical water, J. Supercrit. Fluids 3 (1990)
228232.
[36] J. Schanzenbcher, J.D. Taylor, J.W. Tester, Ethanol oxidation and hydrolysis rates
in supercritical water, J. Supercrit. Fluids 22 (2002) 139147.
[37] L. Ott, V. Lehr, S. Urfels, M. Bicker, H. Vogel, Inuence of salts on the dehydration
of several biomass-derived polyols in sub- and supercritical water, J. Supercrit.
Fluids 38 (2006) 8093.
[38] W. Bhler, E. Dinjus, H.J. Ederer, A. Kruse, C. Mas, Ionic reactions and pyrolysis
of glycerol as competing reaction pathways in near- and supercritical water, J.
Supercrit. Fluids 22 (2002) 3753.
[39] P. Krammer, H. Vogel, Hydrolysis of esters in subcritical and supercritical water,
J. Supercrit. Fluids 16 (2000) 189206.
[40] P.H.L. Moquin, F. Temelli, Kinetic modeling of hydrolysis of canola oil in supercritical media, J. Supercrit. Fluids 45 (2008) 94101.
[41] J.M.L. Penninger, R.J.A. Kersten, H.C.L. Baur, Reactions of diphenylether in supercritical watermechanism and kinetics, J. Supercrit. Fluids 16 (1999) 119132.
[42] J.M.L. Penninger, R.J.A. Kersten, H.C.L. Baur, Hydrolysis of diphenylether in
supercritical water. Effects of dissolved NaCl, J. Supercrit. Fluids 17 (2000)
215226.

381

[43] M. Bicker, S. Endres, L. Ott, H. Vogel, Catalytic conversion of carbohydrates in


subcritical water: a new chemical process for lactic acid production, J. Mol.
Catal. A: Chem. 239 (2005) 151157.
[44] M. Bicker, D. Kaiser, L. Ott, H. Vogel, Dehydration of d-fructose to hydroxymethylfurfural in sub- and supercritical uids, J. Supercrit. Fluids 36 (2005)
118126.
[45] S.D. Iyer, M.T. Klein, Effect of pressure on the rate of butyronitrile hydrolysis in
high-temperature water, J. Supercrit. Fluids 10 (1997) 191200.
[46] C.L. Harrell, J.S. Moscariello, M.T. Klein, The absence of wall effects during benzonitrile hydrolysis, J. Supercrit. Fluids 14 (1999) 219224.
[47] P.A. Marrone, P.M. Gschwend, K.C. Swallow, W.A. Peters, J.W. Tester, Product distribution and reaction pathways for methylene chloride hydrolysis
and oxidation under hydrothermal conditions, J. Supercrit. Fluids 12 (1998)
239254.
[48] J.W. Tester, P.A. Marrone, M.M. DiPippo, K. Sako, M.T. Reagan, T. Arias, W.A.
Peters, Chemical reactions and phase equilibria of model halocarbons and salts
in sub- and supercritical water (200300 bar, 100600 C), J. Supercrit. Fluids
13 (1998) 225240.
[49] A. Kruse, E. Dinjus, Review: hot compressed water as reaction medium and
reactant. Properties and synthesis reactions, J. Supercrit. Fluids 39 (2007)
362380.
[50] A. Kruse, E. Dinjus, Hot compressed water as reaction medium and reactant. 2.
Degradation reactions, J. Supercrit. Fluids 41 (2007) 361379.
[51] H.J. Ederer, A. Kruse, C. Mas, K.H. Ebert, Modelling of the pyrolysis of tertbutylbenzene in supercritical water, J. Supercrit. Fluids 15 (1999) 191204.
[52] D. Mateos, J.R. Portela, J. Mercadier, F. Marias, Ch. Marraud, F. Cansell, New
approach for kinetic parameters determination for hydrothermal oxidation
reaction, J. Supercrit. Fluids 34 (2005) 6370.
[53] M. Osada, M. Watanabe, K. Sue, T. Adschiri, K. Arai, Water density dependence
of formaldehyde reaction in supercritical water, J. Supercrit. Fluids 28 (2004)
219224.
[54] J.A. Onwudili, P.T. Williams, Reaction mechanisms for the decomposition of
phenanthrene and naphthalene under hydrothermal conditions, J. Supercrit.
Fluids 39 (2007) 399408.
[55] Y. Futamura, K. Yahara, K. Yamamoto, Evidence for the production of uorescent pyrazine derivatives using supercritical water, J. Supercrit. Fluids 41 (2007)
279284.
[56] D. Klingler, J. Berg, H. Vogel, Hydrothermal reactions of alanine and glycine in
sub- and supercritical water, J. Supercrit. Fluids 43 (2007) 112119.
[57] K.M. Benjamin, Ph.E. Savage, Hydrothermal reactions of methylamine, J. Supercrit. Fluids 31 (2004) 301311.
[58] V.I. Anikeev, A. Yermakova, V.A. Semikolenov, M. Goto, Effect of supercritical
water density on the rate constant of aliphatic nitrocompounds decomposition,
J. Supercrit. Fluids 33 (2005) 243246.
[59] R. Lachance, I. Paschkewitz, I. DiNaro, I.W. Tester, Thiodiglycol hydrolysis and
oxidation in sub-, and supercritical water, J. Supercrit. Fluids 16 (1999) 133147.
[60] J.B. Dunn, M.L. Burns, S.E. Hunter, Ph.E. Savage, Hydrothermal stability of aromatic carboxylic acids, J. Supercrit. Fluids 27 (2003) 263274.
[61] A.J. Belsky, T.-J. Li, T.B. Brill, Reactions of cyanamide, dicyandiamide and
related cyclic azines in high temperature water, J. Supercrit. Fluids 10 (1997)
201208.
[62] Y. Takeshita, K. Kato, K. Takahashi, Y. Sato, S. Nishi, Basic study on treatment of
waste polyvinyl chloride plastics by hydrothermal decomposition in subcritical
and supercritical regions, J. Supercrit. Fluids 31 (2004) 185193.
[63] Z. Fang, R.L. Smith Jr., H. Inomata, K. Arai, Phase behavior and reaction of
polyethylene terephthalatewater systems at pressures up to 173 MPa and
temperatures up to 490 C, J. Supercrit. Fluids 15 (1999) 229243.
[64] Z. Fang, R.L. Smith Jr., H. Inomata, K. Arai, Phase behavior and reaction of
polyethylene in supercritical water at pressures up to 2.6 GPa and temperatures
up to 670 C, J. Supercrit. Fluids 16 (2000) 207216.
[65] M. Watanabe, H. Hirakoso, S. Sawamoto, T. Adschiri, K. Arai, Polyethylene conversion in supercritical water, J. Supercrit. Fluids 13 (1998) 247252.

Vous aimerez peut-être aussi